Next Article in Journal
A Pilot Study of a Panel of Ocular Inflammation Biomarkers in Patients with Primary Sjögren’s Syndrome
Next Article in Special Issue
FtbZIP85 Is Involved in the Accumulation of Proanthocyanidin by Regulating the Transcription of FtDFR in Tartary Buckwheat
Previous Article in Journal
Distribution of Exonic Variants in Glycogen Synthesis and Catabolism Genes in Late Onset Pompe Disease (LOPD)
Previous Article in Special Issue
Genome-Wide Identification of Sweet Orange WRKY Transcription Factors and Analysis of Their Expression in Response to Infection by Penicillium digitatum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in the Research on Plant WRKY Transcription Factors Responsive to External Stresses

1
College of Ecology, Shanghai Institute of Technology, Shanghai 201418, China
2
Beijing Engineering Research Center of Functional Floriculture, Institute of Grassland, Flowers and Ecology, Beijing Academy of Agriculture and Forestry Sciences, Beijing 100097, China
*
Authors to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2023, 45(4), 2861-2880; https://doi.org/10.3390/cimb45040187
Submission received: 1 March 2023 / Revised: 20 March 2023 / Accepted: 23 March 2023 / Published: 1 April 2023
(This article belongs to the Special Issue Stress and Signal Transduction in Plants)

Abstract

:
The WRKY transcription factors are a class of transcriptional regulators that are ubiquitous in plants, wherein they play key roles in various physiological activities, including responses to stress. Specifically, WRKY transcription factors mediate plant responses to biotic and abiotic stresses through the binding of their conserved domain to the W-box element of the target gene promoter and the subsequent activation or inhibition of transcription (self-regulation or cross-regulation). In this review, the progress in the research on the regulatory effects of WRKY transcription factors on plant responses to external stresses is summarized, with a particular focus on the structural characteristics, classifications, biological functions, effects on plant secondary metabolism, regulatory networks, and other aspects of WRKY transcription factors. Future research and prospects in this field are also proposed.

1. Introduction

Plants are often subjected to a variety of biotic and abiotic stresses, including drought, extreme temperatures, salinity, and other adverse environmental conditions [1]. Therefore, improving the resistance and tolerance of crops to these stresses is an important objective of breeding programs. Plants have evolved many complex molecular, cellular, physiological, and biochemical mechanisms that allow them to cope with abiotic stresses [2,3], including stomatal movement, signal perception and transduction, stress-induced gene expression, and metabolic changes [4]. When plants grow under stress conditions, transcription factors bind to specific cis-acting elements in the promoter region of target genes to regulate expression, signal transduction, and adaptation-related networks [5]. To date, diverse transcription factor families in plants, including WRKY, MYB, NAC, and bHLH, have been identified and functionally characterized regarding their involvement in biotic and abiotic stress responses [6,7,8,9].
The WRKY family of transcription factors is one of the biggest families of transcriptional regulators in plants. These plant-specific transcription factors make up an important part of the signaling network involved in the regulation of a variety of biological processes [10]. The WRKY transcription factors were named because they contain a conserved domain comprising approximately 60 amino acids as well as a conserved WRKYGQK heptapeptide sequence at the N-terminus [11]. In 1994, the first WRKY transcription factor (SPF1) was reported to mediate sucrose-regulated gene expression in sweet potatoes [12]. A subsequent study identified WRKY1, WRKY2, and WRKY3 in parsley (Petroselinum crispum) as additional WRKY transcription factors and revealed for the first time the importance of these types of transcription factors for regulating plant responses to pathogens [13]. The WRKY transcription factors in various plant species have been studied, including Arabidopsis thaliana, rice, tobacco, oat, cucumber, and moss. Moreover, genes encoding WRKY-like proteins have been detected in protozoa (Giardia lamblia and Dictyostelium discoideum) and green algae (e.g., Chlamydomonas reinhardtii), indicative of the ancient origin of the WRKY gene family [14]. Earlier studies confirmed that WRKY transcription factors contribute to plant growth and development and various biotic and abiotic stress responses. Therefore, the regulatory functions and networks of WRKY transcription factors have been a major focus of the research conducted by plant molecular biologists. In this review article, the roles of WRKY transcription factors in plant biotic and abiotic stress responses, plant nutrient stress responses, and plant secondary metabolism are described, while the regulatory mechanisms and networks associated with WRKY transcription factors are summarized.

2. Structural Features and Classification of the WRKY Transcription Factors

The WRKY transcription factors are key regulatory proteins that respond to biotic and abiotic stresses and regulate physiological processes and development [15]. The WRKY family, which was originally identified in sweet potatoes, consists of proteins that contain either one or two WRKY domains. The WRKY domain forms four β-sheets and consists of a conserved WRKYGQK sequence motif as well as a zinc finger (C2H2 or C2HC) structure [16]. These transcription factors regulate the transcription of the downstream target genes by binding specifically to the W-box [(T)TGAC(C/T)] cis-element in the gene promoter [17]. According to the number of WRKY domains and zinc-finger motifs, WRKY transcription factors have been divided into Groups I, II, and III. [17,18]. Specifically, the proteins with two WRKY domains and a C-terminal domain that plays an auxiliary role in the binding to DNA belong to Group I. The proteins in Group II and Group III contain only one WRKY domain and are distinguished by the presence of C2H2 and C2HC motifs, respectively [18]. Group II (i.e., the largest group of WRKY transcription factors) was further subdivided into five subfamilies (a–e) on the basis of the diversity in the conserved structural motifs other than the WRKY domain [19,20]. Group III WRKY transcription factors are associated with plant responses to a variety of diseases and insect pests and are exclusive to higher plants [20] (Figure 1). This is also confirmed by the phylogenetic analysis of the model plant A. thaliana (Figure 2). The WRKY proteins also contain potential basic nuclear localization signals, leucine zippers, as well as serine–threonine-rich, glutamine-rich, and proline-rich regions. The differences in the transcriptional regulatory functions of transcription factors may be due to specific structures (e.g., kinase region) [21].

3. Biological Functions of WRKY Transcription Factors in Plants

Recent comprehensive studies on WRKY family members have shown that these transcription factors are involved in a variety of processes related to biotic and abiotic stress responses (e.g., hormone signaling) [22] while also regulating the expression of downstream genes to promote or inhibit the synthesis of the related proteins [23]. Many WRKY transcription factor family members have been isolated from a variety of plants, including 74 in A. thaliana, 68 in sorghum, and more than 100 in rice, soybean, and other higher plants [10].

3.1. Regulatory Roles in Plant Biotic Stress Responses

Infestations by herbivorous insects and pathogen infections are among the main biotic stresses encountered by plants. During long-term evolution, plants formed two defense mechanisms that protect against biotic stresses. The first mechanism, which affects plant defenses against pathogen infections, mediates the restriction of the pathogen to the damaged site to prevent the infection from spreading (i.e., systemic acquired resistance). The second mechanism, which is induced by external stressors, leads to signaling molecules rapidly activating salicylic acid (SA) pathways. Signal transduction pathways, such as those associated with jasmonic acid (JA) and ethylene (ET), modulate the transcription of related genes and modify proteins, leading to the resistance to various biotic stresses [24,25,26]. Plants produce a suite of preformed secondary metabolites and antibacterial compounds that have inhibitory effects on potential pathogens, but they also possess the following two active defense mechanisms: pathogen-associated molecular-pattern (PAMP)-triggered immunity (PTI) and effector-triggered immunity (ETI) [27,28,29,30] (Table 1).
Shinde et al. reported that WRKY1 positively regulates the resistance of wild tomato (Solanum arcanum) to early blight caused by Alternaria solani. Another study indicated SaWRKY1 clearly up-regulates the expression of XTH5 and MYB2, which influences the disease phenotype [31]. Moreover, the expression levels of 10 WRKY genes, including PnWRKY9, increase in the root. The overexpression of PnWRKY9 in transgenic tobacco plants results in a significant increase in the resistance to F. solani, while the RNAi-mediated decrease in PnWRKY9 expression increases the susceptibility of P. notoginseng leaves to F. solani. Furthermore, PnWRKY9 and the JA signaling pathway synergistically affect the disease resistance of transgenic tobacco [32]. The overexpression of the rice (Oryza sativa) OsWRKY45 transcription factor gene results in increased resistance to Magnaporthe oryzae [33]. Twenty-five LrWRKY genes were identified during a previous analysis of the Lilium regale transcriptome, including fifteen LrWRKY genes whose expression is significantly induced by Botrytis cinerea. Compared with the wild-type control, transgenic A. thaliana plants overexpressing LrWRKY39 and LrWRKY41a are more resistant and sensitive to B. cinerea, respectively [34]. However, the expression of FvWRKY42, which was isolated from wild strawberry (Fragaria vesca) (Heilongjiang No. 3; diploid), is induced by multiple stresses; the encoded transcription factor interacts with various stress response-related proteins. The overexpression of FvWRKY42 in A. thaliana leads to increased powdery mildew resistance because of the associated increase in fungal cell death, sporulation, and delayed mycelial growth [35]. In Panax notoginseng treated with methyl jasmonate (MeJA) and infected with Fusarium solani, the ectopic expression of AtWRKY18 results in up-regulated PR gene expression and resistance to the bacterial pathogen Pseudomonas syringae [36]. An earlier analysis of the interaction between AkWRKY and the W-box of AkNBS genes in three-leaf akebia (Akebia trifoliata) showed that AkWRKY24 can enhance disease resistance by positively regulating the expression of AkNBS18 [37]. In A. thaliana, more than two-thirds of the identified AtWRKY members are affected by bacterial infections and SA treatments, suggestive of their key roles in biotic stress responses [38,39,40,41].
Earlier studies have shown that WRKY transcription factors can function both upstream and downstream of hormones and contribute to the antagonistic effects between SA and JA/ET, while also controlling developmental processes via auxin, cytokinin, and brassinosteroids [50]. Signaling pathways mediated by SA are associated with the resistance to biotrophic and semi-biotrophic pathogens, whereas JA/ET signaling pathways are typically associated with the resistance to necrotrophic pathogens [33]. In a study on the potential effects of ET during an infection of the A. thaliana EIN3/EIL1 double mutant, Rishmawi et al. observed that WRKY75 is directly regulated by ET. In the presence of ACC, WRKY75 expression is up-regulated at 7 days post-infection. Moreover, ET has dual roles in infected rapeseed plants and contributes to root defenses against pathogens [51]. The overexpression of EjWRKY17 in loquat (Eriobotrya japonica) promotes abscisic-acid (ABA)-mediated stomatal closure under drought conditions and significantly up-regulates the expression of ABA biosynthesis genes and genes related to stress responses [52]. The overexpression of the SA-inducible gene PtrWRKY89 increases PR gene expression and the resistance of transgenic cottonwood trees to pathogens [53]. Most WRKY transcription factors are negative regulators, but a few are known to function as positive regulators. In peanut (Arachis hypogaea) plants, AhWRKY7, AhWRKY8, and AhWRKY13 expression levels are down-regulated in response to SA and JA, which is in contrast to the up-regulated expression of the AhWRKY1 and AhWRKY12 genes that encode transcription factors that modulate SA and JA signal transduction pathways, reflecting their importance for disease resistance [54].

3.2. Regulatory Roles in Plant Abiotic Stress Responses

In addition to various biotic stresses, plants are exposed to abiotic stresses throughout their growth period, including high and low temperatures, salinity, drought, and nutrient deficiency [55]. These stresses, which can occur simultaneously, have detrimental effects on plant physiological and biochemical processes and severely restrict growth and development. The expression of WRKY genes is induced in response to different abiotic stresses, implying they likely influence abiotic stress tolerance. The precise regulation of WRKY proteins during plant stress responses is associated with the establishment of a complex signaling network. Given the crucial functions of WRKY proteins in plant abiotic stress responses, they are probably key contributors to stress tolerance [56,57]. Several recent investigations indicated that manipulating WRKY transcription factor levels in genetically modified plants (i.e., gene knockout or overexpression) alters specific stress responses. Some abiotic stresses strongly and rapidly induce the expression of numerous WRKY-encoding genes (e.g., drought or salinity, low and high temperatures, or osmotic stress) [10] (Table 2).

3.2.1. Temperature Stress

Frequent changes in temperature and extreme weather conditions worldwide result in high- and low-temperature stresses. Thus, temperature stress is one of the main abiotic stresses affecting plants. Temperature extremes alter plant physiological indices and affect plant growth and development, ultimately resulting in economic losses for agriculturally important plant species. Therefore, plant cells must be protected from the detrimental effects of temperature changes. Accordingly, temperature-stress-induced changes in plants should be elucidated to provide the theoretical basis for crop breeding and the development of improved agronomic practices to increase agricultural production [58,59]. Many WRKY transcription factors participate in plant responses to temperature stress by regulating the expression of related genes. For example, the expression of 10 OsWRKY genes in rice is significantly altered by different environmental stresses, including cold/heat, salinity (sodium chloride treatment), and simulated drought (e.g., polyethylene glycol treatment) [60]. Rizhsky et al. determined that most tobacco WRKY gene expression levels are altered by cold/heat stresses [61].
High-temperature stress can damage plant cell membranes and disrupt reactive oxygen species (ROS) homeostasis, resulting in changes to plant structures and functions [62]. Recent research showed that in medicinal dandelion, WRKY gene expression is up-regulated substantially more under high-temperature stress conditions than in response to low-temperature stress [63]. Dang [64] reported that the expression of CaWRKY40 increases significantly in the leaves of pepper (Capsicum annuum) plants grown at high temperatures. In A. thaliana, the expression levels of both AtWRKY25 and AtWRKY26 increase following a high-temperature treatment, whereas the expression of AtWRKY33 decreases; however, mutations to all three genes increase the sensitivity of A. thaliana plants to high-temperature stress [65]. The initial stage of the wheat streak rust infection of wheat is characterized by a significant increase in the TaWRKY70 expression level at elevated temperatures. This finding reflects the positive correlation between TaWRKY70 expression and the resistance of wheat seedlings to heat stress as well as the likely activation of the SA and ET signaling pathways during the initial stage of the infection [66]. Wu et al. isolated a heat-induced differentially expressed lily (Lilium longiflorum) gene belonging to WRKY Group IIe (LlWRKY22). This gene encodes a protein involved in the mechanism underlying heat tolerance. Additional research revealed that the overexpression of LlWRKY22 increases the heat tolerance of transgenic plants and up-regulates the expression of AtDREB2A, AtDREB2B, AtDREB2C, and AtJUB1. However, the LlWRKY22 regulatory pathway is also associated with the ABA signaling pathway [67].
The analysis of the cotton (Gossypium hirsutum) transcriptome identified 10 WRKY genes with up-regulated expression levels under low-temperature stress conditions [68]. Using RNA-seq and chromatin immunoprecipitation sequencing data, Guo et al. showed that SlWRKY33 directly targets and induces the expression of multiple genes encoding kinases, transcription factors, and molecular chaperones, including CDPK11, MYBS3, and BAG6, thereby enhancing cold tolerance [69]. The overexpression of OsWRKY76 in rice (O. sativa) positively affects cold tolerance [70]. Furthermore, AtWRKY34 negatively regulates the sensitivity of mature pollen grains to cold stress by inhibiting the expression of CBF pathway genes [71].

3.2.2. Drought Stress and Salt Stress

Two major environmental factors that significantly limit global crop production are drought and high salinity. More specifically, these two environmental stresses lead to an intracellular ionic and osmotic imbalance while also interfering with photosynthetic activities, cellular energy consumption, and the redox balance. The WRKY family of transcription factors is involved in the regulatory system that links the perception and transduction of environmental signals with adaptive cellular responses. Many WRKY transcription factors were identified as the primary factors regulating molecular programming, which can enhance plant stress resistance. Some WRKY transcription factors are related to the drought and salt stress tolerance of A. thaliana, rice, and other plants. Several established mechanisms that mediate drought and salinity tolerance are affected by ABA signaling. More specifically, ABA has been implicated in plant responses to drought conditions, and components of the ABA signaling pathway in the nucleus are targeted by signals from plastids and mitochondria, which affects the retrograde signal transduction in plant cells [72,73].
Yan et al. detected induced GhWRKY17 expression in plants treated with various stresses (i.e., drought, salt, H2O2, and ABA). On the basis of the germination rate, root growth, survival rate, leaf water loss, and chlorophyll content, constitutive GhWRKY17 expression significantly decreases the tolerance of Nicotiana benthamiana to drought and salt stresses [74]. However, IgWRKY50 and IgWRKY32 in Iris germanica encode positive regulators that enhance the drought resistance of A. thaliana transgenic plants by modulating the ABA signal transduction pathway [75]. The expression of GsWRKY20 in transgenic A. thaliana plants leads to increased drought tolerance [76]. In a separate study, GmWRKY6 overexpression was confirmed to improve the resistance of Lotus japonicus to salt stress, whereas the overexpression of GmWRKY12 enhances the drought resistance and salt tolerance of hairy roots [77]. Many cotton WRKY genes play a crucial regulatory role during the stress responses of model plants. For example, the overexpression of GhWRKY34 and GhWRKY41 reportedly increases the salt tolerance of A. thaliana and the drought tolerance of tobacco [78,79]. Additionally, GhWRKY1 is a transcriptional regulator related to the drought tolerance of upland cotton (G. hirsutum). The overexpression of genes encoding GhWRKY1-like proteins in A. thaliana increases drought tolerance by manipulating the synthesis of ABA and interactions with several cis-elements [80].
Under drought conditions, PmWRKY6, PmWRKY10, and PmWRKY30 expression levels are up-regulated in different plants and tissues, while PmWRKY22 expression is down-regulated. Compared with wild-type tobacco, PmWRKY31-expressing transgenic tobacco plants contain less malondialdehyde but more proline. Significant increases in the expression of the related genes have been detected in PmWRKY31-expressing transgenic tobacco, which leads to an increase in drought tolerance [81]. Moreover, AtWRKY33 is an upstream regulator of AtCYP94B1 in A. thaliana; mutations to AtWRKY33 result in decreased suberin levels and salt-sensitive phenotypes, further confirming that AtWRKY33-mediated AtCYP94B1 regulation is part of the salt tolerance mechanism [82]. The A. thaliana genes AtWRKY21, AtWRKY46, AtWRKY54, and AtWRKY70 encode regulators of osmotic stress responses [83,84]. The DgWRKY1 transcription factor positively regulates the salt stress response [85]. A previous study indicated that salt stress induces PeWRKY1 expression in Populus euphratica and that the positive effect of PeWRKY1 expression on the salt tolerance of transgenic tobacco lines is associated with the promotion of Na+ efflux from the root cells [86].

3.2.3. Nutrient Stress

Normal plant growth and development depends on the uptake of various nutrients, including nitrogen, phosphorus, and potassium. Plant morphogenesis is severely affected by the lack of essential elements. Phosphorus is one of the essential nutrients required by plants, representing approximately 0.2% of the plant dry weight. In plants and animals, phosphorus helps regulate key metabolic pathways as well as many enzymatic reactions. Plants absorb phosphorus in the form of phosphate. The first WRKY transcription factor revealed to be associated with phosphorus deficiency was AtWRKY75. Under phosphorus-deficient conditions, AtWRKY75 expression is strongly induced in A. thaliana. The inhibited expression of this gene increases the sensitivity of plants to phosphorus stress. Furthermore, the uptake of phosphorus by plants decreases in low-phosphorus environments [87]. Among the WRKY genes in cotton (Gossypium barbadense), GbWRKY1 regulates plant responses to phosphorus deficiency. The overexpression of GbWRKY1 in A. thaliana can decrease the severity of the effects of phosphorus deficiency, induce the accumulation of phosphorus, and promote lateral root development and phosphatase activity [88]. Shen et al. identified WRKY33 as a negative regulator of phosphorus deficiency-induced root architecture remodeling in A. thaliana. More specifically, it controls the transcription of ALMT1 under phosphorus-deficient conditions, thereby modulating the accumulation of Fe3+ in the root tips and inhibiting root growth [89].
Nitrogen and potassium are also critical nutrients required by plants. In rice (O. sativa), the expression levels of many WRKY genes are modified (up-regulated or down-regulated) by low-nitrogen stress [90]. Of the wheat (Triticum aestivum) WRKY genes, 10 (e.g., WRKY4, WRKY6, WRKY12, and WRKY18) encode positive regulators of plant responses to potassium deficiency, whereas the other genes encode negative regulators [91]. The WRKY transcription factors are widely implicated in alleviating the deleterious effects of plant nutrient stress, especially nitrogen, phosphorus, and potassium deficiencies. Furthermore, specific micronutrients are important for plant growth and developmental processes. Earlier research showed WRKY transcription factors are involved in altering plant activities in response to a lack of trace elements. For example, under iron stress conditions, WRKY46 can directly regulate the transcription of VITL1 (vacuolar iron transporter1-like1) by specifically binding to the W-box region in the gene promoter, which affects iron uptake and transport in plants [92]. In response to boron stress, the WRKY transcription factors in A. thaliana induce the expression of NIP5;1 (nodulin 26-like intrinsic protein 5;1), allowing plants to adapt to the limited availability of boron, indicating that WRKY6 can regulate root tip gene expression under boron-deficient conditions [93]. Many WRKY family members that regulate plant responses to micronutrient stress have been identified. In wheat (T. aestivum), an exposure to calcium stress up-regulates the expression of four WRKY genes, including WRKY18, but has the opposite effect on WRKY6 expression. Under zinc stress conditions, the expression of six WRKY genes (e.g., WRKY3) is up-regulated in wheat, but the expression of three other WRKY genes (e.g., WRKY17) is down-regulated [91].
The WRKY transcription factors also participate in the regulation of plant responses to other elements, some of which have beneficial effects. In A. thaliana, AtWRKY46 is a negative regulator associated with plant responses to aluminum stress, which inhibits AtWRKY46 expression. If A. thaliana plants are treated with toxic levels of aluminum, WRKY46 binds directly to the W-box of the ALMT1 (aluminum-activated malate transporter 1) promoter, thereby negatively regulating ALMT1 expression and improving the resistance to aluminum stress [94]. The expression of PyWRKY75, which was isolated and cloned from poplar (Populus yunnanensis), significantly promotes cadmium uptake and accumulation in trees, with cadmium levels that are 51.32% higher in PyWRKY75-overexpressing trees than in the wild-type controls. Antioxidants, such as peroxidase (POD), superoxide dismutase (SOD), catalase (CAT), ascorbate peroxidase (APX), ascorbic acid (AsA), reduced glutathione (GSH), and phytochelatins (PCs), can increase the cadmium stress tolerance of transgenic poplar trees and accumulate differentially in the roots, shoots, and leaves [95]. These findings provide important insights into the contributions of WRKY transcription factors to plant responses to nutrient stress.
Table 2. Regulation of WRKY transcription factors in abiotic stress responses.
Table 2. Regulation of WRKY transcription factors in abiotic stress responses.
Type of StressSpeciesGeneReferences
High temperaturePepper (Capsicum annuum)CaWRKY40[64]
Arabidopsis (Arabidopsis thaliana)AtWRKY25, AtWRKY26,
AtWRKY33
[65]
Wheat (Triticum aestivum)TaWRKY70[66]
Lily (Lilium browniivar)LlWRKY22[67]
Low temperatureCultivated tomato (Solanum lycopersicum)SlWRKY33[69]
Rice (Oryza sativa)OsWRKY76[70]
Arabidopsis (Arabidopsis thaliana)AtWRKY34[71]
DroughtGerman iris (Iris germanica)IgWRKY50, IgWRKY32[75]
Wild soybean (Glycine soja)GsWRKY20[76]
Soybean (Glycine max)GmWRKY12[77]
Upland cotton (Gossypium hirsutum)GhWRKY1[80]
Pinus massoniana (Pinus massoniana)PmWRKY31[81]
SaltSoybean (Glycine max)GmWRKY6[77]
Cotton (Gossypium hirsutum)GhWRKY34, GhWRKY41[78,79]
Arabidopsis (Arabidopsis thaliana)AtWRKY33[82]
Chrysanthemum (Dendranthema grandiflorum)DgWRKY1[85]
Diversiform-leaved poplar (Populus euphratica)PeWRKY1[86]
PhosphorusArabidopsis (Arabidopsis thaliana)AtWRKY75[87]
Gossypium barbadense (Gossypium barbadense)GbWRKY1[88]
AluminumArabidopsis (Arabidopsis thaliana)AtWRKY46[93]
Soybean (Glycine max)GmWRKY21[94]
CadmiumPoplar (Populus yunnanensis)PyWRKY75[95]
IronRice (Oryza sativa)OsWRKY74[96]

3.2.4. Oxidative Stress

Of the other stresses that affect plants, oxidative stress is one of the most harmful [97]. In plants, the four major types of ROS are singlet oxygen (O2), hydroxyl radicals (OH·), superoxide anions (O2), and hydrogen peroxide (H2O2). Therefore, the active oxygen signaling network modulated by WRKY transcription factors has important biological functions in plants (Table 3).
Various WRKY transcription factors in A. thaliana are activated by H2O2 stress. Cytoplasmic ascorbate peroxidase 1 (Apx1) is an H2O2 scavenger. An earlier microarray analysis of Apx1-deficient A. thaliana plants by Rizhsky et al. revealed that the production of two zinc-finger proteins (Zat12 and Zat7) and WRKY25 increases in Apx1-knockout plants cultured under controlled conditions. Thus, Zat12 plays an important role in the signal transduction network mediating the response of A. thaliana to oxidative stress. Specifically, Zat12 is required for the expression of Zat7, WRKY25, and Apx1 under oxidative stress conditions [97].
Studies have shown that the adaptation of A. thaliana to underwater hypoxic conditions involves the interaction between WRKY33 and WRKY12, which positively regulates the expression of the ethylene-responsive factor VII gene RAP2.2; the overexpression of WRKY33 and WRKY12 increases plant resistance to hypoxia [98]. Another study demonstrated that WRKY53 can interact directly with a mitogen-activated protein kinase (MEKK1) and regulate the expression of CAT1, CAT2, and CAT3, which help mediate antioxidant defenses [99]. The activated expression of AtWRKY53 inhibits stomatal closure by decreasing the H2O2 content and induces stomatal opening by promoting starch degradation [100]. The overexpression of GhWRKY68 impairs the ability of plants to tolerate oxidative stress under both drought and saline conditions [101]. An earlier study by Yan et al. compared ROS levels in GhWRKY68-overexpressing transgenic and wild-type plants under normal and stressed conditions. Under normal conditions, both H2O2 and O2 accumulated at low levels, and there were no significant differences between the wild-type and transgenic lines; however, after the exposure to drought and saline conditions, the transgenic lines accumulated more H2O2 and O2, implying that GhWRKY17 increased ROS levels after the exposure to drought and salt stresses [74]. Cold and excessive salinity can substantially induce VvWRKY28 expression in grapes (Vitis vinifera). The expression of VvWRKY28 can considerably increase the tolerance of A. thaliana to low temperatures and high salinity, likely because of the associated increase in SOD, POD, and CAT activities [102]. Therefore, WRKY transcription factors have important biological functions influencing the active oxygen signaling network in plants.

4. Resistance-Related Regulatory Effects of WRKY Transcription Factors on Plant

Secondary Metabolism

The role of WRKY transcription factors during the biosynthesis of natural products [103] has recently become a topic of interest among researchers. The WRKY family of transcription factors is involved in plant metabolic processes that result in the production of many secondary metabolites, including lignin, plant hormones, phenols, alkaloids, terpenes, and flavonoids. Secondary metabolites have crucial functions affecting plant growth, development, and responses to environmental stimuli (Table 4).
In Isatis indigotica, IiWRKY34 expression levels increase following drought and salt treatments, with IiWRKY34 subsequently binding to the W-box in the promoter region of Ii4CL3, which encodes the key rate-limiting enzyme for lignin biosynthesis. The accumulation of lignin increases the biological activity and stress resistance of plants [104]. Maize (Zea mays) plants accumulate terpenoid phytoalexins, liquiritigenin, and zeatin in response to various stimuli. The biosynthesis of maize terpenoid phytoalexins is regulated by ZmWRKY79, which is a potential major regulator of stress responses through its effects on plant hormone metabolism or signal transduction and ROS scavenging [105].
Research on the regulatory effects of WRKY transcription factors on alkaloid synthesis has mostly focused on the indole alkaloid biosynthesis pathway. The CjWRKY1 transcription factor in Coptis japonica was the first transcription factor confirmed to regulate alkaloid production. More specifically, it affects the biosynthesis of benzylisoquinoline alkaloids [106]. Moreover, CrWRKY1 controls the biosynthesis of terpene indole alkaloids, which have anti-tumor effects, by regulating the expression of the tryptophan decarboxylase (TDC)-encoding gene in Catharanthus roseus [107]. In Ophiorrhiza pumila, OpWRKY6 negatively regulates the biosynthesis of camptothecin in the iridoid pathway and the shikimate pathway by directly down-regulating the expression of OpGES, Op10HGO, Op7DLH, and OpTDC [108].
An earlier analysis of terpenoid biosynthesis in tomato (Solanum lycopersicum) revealed that SlWRKY71 regulates the expression of the terpenoid synthase gene following a JA treatment [109]. Jasmonide is a diterpenoid plant antitoxin with allelopathic effects and is widely distributed in feathery bryophytes. Examinations of cis-acting elements and binding sites indicated that stress induces CpWRKY expression, which influences jasmonide biosynthesis [104]. Artemisinin is currently the most effective substance for treating malaria. The AaWRKY1 gene, which was isolated from secretory glandular trichomes, encodes a protein that binds to the W-box of the ADS promoter to activate transcription and promote artemisinin biosynthesis [110]. The cDNA sequences of many WRKY genes in Gossypium arboreum have been isolated. One of these genes, GaWRKY1, encodes a protein containing a single WRKY domain and a putative leucine zipper at the N-terminus. In suspension cells, fungal elicitors and MeJA induce GaWRKY1 and CAD1-A expression and sesquiterpene aldehyde biosynthesis [111].
In addition, some WRKY transcription factors can negatively regulate flavonoid synthesis in plants. For example, the heterologous expression of the Brassica napus gene BnWRKY41-1 in A. thaliana rosette leaves reportedly leads to decreases in the anthocyanin content [112]. Long-term studies on rice (O. sativa) have shown that OsWRKY13 up-regulates CHS expression and controls flavonoid phytoalexin biosynthesis. Furthermore, WRKY proteins may be important for increasing the expression of OsWRKY13 and genes involved in defense responses to pathogens [113]. Recent research revealed a new disease defense mechanism, the WRKY–MAPK pathway, which protects plants against pathogens by promoting flavonoid biosynthesis [114].

5. WRKY Transcription Factor Regulatory Network

At the DNA level, each WRKY transcription factor recognizes and binds to the W-box domain in its target genes, which may include its own gene, to activate or inhibit transcription. Interactions between WRKY transcription factors and the downstream target genes give rise to a complex WRKY regulatory network. At the protein level, WRKY transcription factors and diverse regulatory proteins coordinately respond to a variety of environmental stresses (Figure 3).

5.1. Self-Regulation and Cross-Regulation by WRKY Transcription Factors

The promoters of WRKY transcription factor target genes and the promoters of most WRKY transcription factor genes contain W-box elements. The WRKY transcription factors control various stress responses via self-regulation as well as the cross-regulation of other WRKY transcription factors [114,115]. For example, the promoter region of PcWRKY1 has multiple W-box elements that mediate the binding of WRKY transcription factors [116]. In addition to binding to the W-box in its own promoter, PcWRKY1 can also bind to the W-box in the PcWRKY3 promoter region [117]. Wu et al. recently identified LlWRKY22 as a novel regulator of lily plant responses to high-temperature stress. Specifically, it induces the expression of its own gene to form a positive feedback loop, while also activating the expression of LlDREB2B, which encodes a core regulator of heat stress responses [67]. In P. euphratica, PeWRKY1 binds to the W-box of the PeHA1 and PeMAX2 promoters to regulate expression. Therefore, PeWRKY1 improves ion homeostasis in P. euphratica in two ways [76]. In banana (Musa acuminata), MaWRKY21 binds directly to the W-box of the MaICS promoter and negatively regulates transcription, thereby decreasing enzyme activity levels [118].

5.2. WRKY Transcription Factors Interact with the Downstream Target Genes

The WRKY transcription factors bind specifically to the W-box (TTGACC/T) of the downstream target genes via their WRKY domain [10,119,120]. For example, under disease stress conditions, WRKY transcription factors bind to the W-box of the tobacco chitinase gene TDBA12 [121]. In A. thaliana, AtWRKY8 mediates the resistance to Phytophthora infestans by interacting with the genes downstream of the MAPKKKα–MEK2–WIPK signaling cascade, which results in H2O2 accumulation and the apoptosis of plant cells [122]. In P. euphratica, PeWRKY1 binds to the W-box element in the promoter of PeHA1 (encoding a plasma membrane H+-ATPase) to increase transcription. The PeWRKY1–PeHA1/PeMAX2–PeGRP2 signaling network that maintains the ionic balance in P. euphratica provides important insights into the molecular basis of the P. euphratica response to salinity stress [87]. Furthermore, AtWRKY6, AtWRKY42, and AtWRKY72 control plant responses to nutrient stress (e.g., low phosphorus levels) by regulating the expression of the downstream target genes [123,124]. The W-box is a cis-acting element that is generally associated with up-regulated expression, but the binding of a WRKY transcription factor to the W-box in the AtWRKY18 promoter in A. thaliana results in down-regulated expression [47]. The WRKY8 transcription factor regulates the expression of the downstream genes, such as RD29A, after interacting with the VQ9 protein [125].

5.3. WRKY Transcription Factors Involved in Plant Hormone Signal Transduction

The WRKY transcription factors are critical components of the plant signaling network that regulates growth and development under biotic and abiotic stress conditions. These transcription factors may function as activators or repressors in a transcription factor network modulating the transduction of signals from organelles and the cytoplasm to the nucleus. They may have roles upstream and downstream of hormones; contribute to the antagonism among the SA, JA, and ET pathways; and control development-related processes through auxin, cytokinin, and brassinosteroids. The SA-mediated signaling pathway is generally associated with the resistance to biotrophic and semi-biotrophic pathogens, whereas JA/ET signaling pathways are associated with the resistance to necrotrophic pathogens [32,33,34]. For example, WRKY70 activates the expression of SA-responsive genes but has the opposite effect on JA-responsive genes. Thus, it integrates the signals from these antagonistic pathways [126]. He et al. [127] determined that the synergistic effects of OsWRKY51 and OsWRKY71 inhibit gibberellin signaling in rice (O. sativa) seed aleurone cells.
Cytokinins and ABA are key hormones influencing plant growth, development, and stress responses. In rose (Rosa hybrida), RhWRKY13 binds to the promoters of the cytokinin oxidase/dehydrogenase 3 (RhCKX3) gene and the ABA-responsive ABA INSENSITIVE4 (RhABI4) gene, leading to the simultaneous inhibition of gene expression in rose petals [128]. Increased cytokinin contents and inhibited ABA responses contribute to the protection of plants from B. cinerea infections. The WRKY2, WRKY18, WRKY40, WRKY60, and WRKY63 transcription factors reportedly modulate the expression of genes encoding ABA response element-binding factors (ABFs/AREBs) by binding to the W-box sequence of the corresponding gene promoters. The ABA signal in the nucleus is a target of plastid and mitochondrial signals which influence retrograde signaling in plant cells [39,129,130]. Studies have shown that WRKY15 is involved in the transduction of signals from the mitochondria to the nucleus and that auxin and cytokinin control plant development with the assistance of WRKY transcription factors [131]. The A. thaliana WRKY57 transcription factor can interact with the JA signaling pathway inhibitors JAZ4/JAZ8 and the auxin signaling pathway inhibitor IAA29, which links the JA- and auxin-mediated plant leaf senescence signaling pathways. Furthermore, WRKY57 binds directly to the promoters of the ABA-related genes RD29A and NCED3 to regulate plant drought resistance [129]. In A. thaliana, WRKY8 aids in the regulation of antiviral responses by regulating ABA and ET signal transduction pathways [125]. Previous research indicated that GbWRKY1 influences cotton resistance to Verticillium wilt via the SA, JA, and ET signaling pathways. Subsequent investigations suggested that GbWRKY1 mainly regulates plant growth and development by participating in ET and other signaling pathways and affecting cytoprotective enzyme activities and resistance-protein-coding gene transcription to improve plant resistance to Verticillium wilt [88].

5.4. Other Regulatory Networks

In addition to the W-box element, WRKY transcription factors can also bind to other sequences. For example, the aging-related transcription factor AtWRKY53 must interact with MEKK1, a MAP kinase, before it can bind to the W-box element [132]. In barley (Hordeum vulgare), WRKY transcription factors related to glucose metabolism can bind to SURE (sugar responsive) elements to facilitate starch synthesis [133]. Zhang et al. identified 104 WRKY33 target genes in A. thaliana under waterlogging stress conditions. A novel cis-element (TCTCTC) designated as the TC box is the major motif in the target gene promoters. This cis-element differs from the W-box previously identified to interact with WRKY33. Genes with a promoter containing this element may be regulated by WRKY33 under flooding conditions. This suggests that the WRKY33 function during the plant response to flooding conditions may depend on the TC box in the promoters of the downstream target genes [134]. The PnWRKY9 recombinant protein overexpressed in tobacco (Nicotiana tabacum) was observed to bind specifically to the W-box sequence in the promoter of a JA-responsive and fungal-resistance-related defensin gene (PnDEFL1) in eggplant. According to yeast one-hybrid assays, PnWRKY9 can activate PnDEFL1 transcription; a tobacco co-expression experiment involving β-glucosidase as a reporter verified these findings [32].

6. Summary and Future Outlook

Much of the research on WRKY transcription factors conducted to date has involved model plants, including arabidopsis, tobacco, rice, soybean, and other economic crops. However, the role of many WRKY transcription factors during evolution and their functional differences must be more thoroughly characterized to further improve plant stress resistance and optimize the breeding of new varieties. Although their biological functions have been extensively studied, there are still some gene functions and modes of action that are unclear. Whole-genome and transcriptome analyses will help to elucidate the functions of WRKY transcription factors during plant stress responses. Molecular biology-based techniques should be applied to mine genomic data, functionally annotate WRKY transcription factors, and clarify the synergistic response mechanisms of other transcription factors.
WRKY transcription factors in the calcium ion (Ca2+) signal transduction pathway associated with plant–pathogen interactions induce the expression of defense-related genes, implying it may be important for plant disease resistance [135,136]. However, the plant response and gene regulatory system mediated by the calcium ion (Ca2+) signal transduction pathway associated with WRKY transcription factors and plant-pathogen interactions are not well understood. Future studies should aim to integrate sequencing data with information regarding protein–DNA and protein–protein interactions to more precisely characterize the regulation of important biological functions by WRKY tran-scription factors, which may lead to the construction of a comprehensive regulatory network. A thorough understanding of the mechanisms underlying WRKY transcription factor functions at the molecular level combined with molecular marker-assisted breeding and biotechnological tools may enable researchers and breeders to develop and select enhanced varieties of economically important crops.

Author Contributions

Conceptualization, X.C.; Data curation, X.C. and H.W.; Formal analysis, H.W. and D.Y.; Investigation, X.C., D.Y. and H.W.; Project administration, D.C. and C.L.; Resources, X.C., H.L. and D.C.; Supervision, X.C. and C.H.; Writing—original draft, H.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grants from the National Natural Science Foundation of China (Grant No. 32102412) and the Beijing Innovation Consortium of Agriculture Research System (BAIC09-2023).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, C.S.; Hou, N.; Fang, N.; He, J.Q.; Ma, Z.Q.; Guan, Q.M.; Li, X.W. Cold shock protein 3 plays a negative role in apple drought tolerance by regulating oxidative stress response. J. Plant Physiol. Biochem. 2021, 168, 83–92. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, Y.; Zhou, Y.; Zhang, D.; Tang, X.; Li, Z.; Shen, C.; Han, X.; Deng, W.; Yin, W.; Xia, X. PtrWRKY75 overexpression reduces the stomatal aperture and improves drought tolerance by salicylic acid-induced reactive oxygen species accumulation in poplar. Environ. Exp. Bot. 2020, 176, 104117. [Google Scholar] [CrossRef]
  3. Cocucci, S.; Morgutti, S.; Cocucci, M.; Scienza, A. A possible relationship between stalk necrosis and membrane transport in grapevine cultivars. J. Sci. Hortic. 1988, 34, 67–74. [Google Scholar] [CrossRef]
  4. Zhang, K.; Wang, Y.; Sun, W.M.; Han, K.M.; Yang, M.; Si, Z.Z.; Li, G.L.; Qiao, Y.K. Effects of exogenous salicylic acid on the resistance response of wild soybean plants (Glycine soja) infected with Soybean mosaic virus. J. Can. J. Plant Pathol. 2020, 42, 84–93. [Google Scholar] [CrossRef]
  5. Zhu, H.; Zhou, Y.; Zhai, H.; He, S.; Zhao, N.; Liu, Q. A novel sweetpotato WRKY transcription factor, IbWRKY2, positively regulates drought and salt tolerance in transgenic Arabidopsis. Biomolecules 2020, 10, 506. [Google Scholar] [CrossRef] [PubMed]
  6. Lv, B.; Wu, Q.; Wang, A.; Li, Q.; Dong, Q.; Yang, J.; Zhao, H.; Wang, X.; Chen, H.; Li, C. A WRKY transcription factor, FtWRKY46, from Tartary buckwheat improves salt tolerance in transgenic Arabidopsis thaliana. Plant Physiol. Biochem. 2020, 147, 43–53. [Google Scholar] [CrossRef]
  7. Gullì, M.; De, P.M.; Perrotta, C.; Rampion, P. A stress-related transcription factor belonging to the YL-1 family is differently regulated in durum wheat cultivars differing in drought sensitivity. J. Plant Physiol. Biochem. 2022, 170, 307–315. [Google Scholar] [CrossRef]
  8. Chen, M.; Na, S. Overexpression of maize MYB-IF35 increases chilling tolerance in Arabidopsis. J. Plant Physiol. Biochem. 2019, 135, 167–173. [Google Scholar]
  9. Xuan, C.N.; Sun, H.K.; Hussain, J.A.; Yeji, Y.; Hay, J.H.; Ju, S.Y.; Chae, O.L.; Yun, D.J.; Woo, S.C. A positive transcription factor in osmotic stress tolerance, ZAT10, is regulated by MAP kinases in Arabidopsis. J. Plant Biol. 2016, 59, 55–61. [Google Scholar]
  10. Rushton, P.J.; Somssich, I.E.; Ringler, P.; Shen, Q.J. WRKY transcription factors. Trends Plant Sci. 2010, 15, 247–258. [Google Scholar] [CrossRef]
  11. Shabir, H.; Wani, S.A.; Balwant, S.; Abhishek, B.; Rohit, J. WRKY transcription factors and plant defense responses: Latest discoveries and future prospects. J. Plant Cell Rep. 2021, 40, 1–15. [Google Scholar]
  12. Ishiguro, S.; Nakamura, K. Characterization of a cDNA encoding a novel DNA-binding protein, Spf1, that recognizes Sp8 sequences in the 5 upstream regions of genes coding for sporamin and beta-amylase from sweet potato. J. Mol. Gen. Genet. 1994, 244, 563–571. [Google Scholar] [CrossRef] [PubMed]
  13. Rushton, P.J.; Macdonald, H.; Huttly, A.K.; Lazarus, C.M.; Hooley, R. Members of a new family of DNA-binding proteins bind to a conserved Cis-element in the promoters of A-Amy2 genes. J. Plant Mole Biol. 1995, 29, 691–702. [Google Scholar] [CrossRef] [PubMed]
  14. Madhunita, B.; Ralf, O. WRKY transcription factors: Jack of many trades in plant. J. Plant Signal. Behav. 2014, 9, e27700. [Google Scholar]
  15. Rushton, D.L.; Tripathi, P.; Rabara, R.C.; Lin, J.; Ringler, P.; Boken, A.K.; Langum, T.J.; Smidt, L.; Boomsma, D.D.; Emme, N.J.; et al. WRKY transcription factors: Key components in abscisic acid signaling. J. Plant Biotechnol. 2012, 10, 2–11. [Google Scholar] [CrossRef] [PubMed]
  16. Bao, W.Q.; Wang, X.W.; Chen, M.; Chai, T.Y.; Wang, H. A WRKY transcription factor, PcWRKY33, from Polygonum cuspidatum reduces salt tolerance in transgenic Arabidopsis thaliana. Plant Cell Rep. 2018, 37, 1033–1048. [Google Scholar] [CrossRef]
  17. Wang, Y.J.; Jiang, L.; Chen, J.Q.; Tao, L.; An, Y.M.; Cai, H.S.; Guo, C.H. Overexpression of the alfalfa WRKY11 gene enhances salt tolerance in soybean. PLoS ONE 2018, 13, e0192382. [Google Scholar] [CrossRef]
  18. Li, W.X.; Pang, S.Y.; Lu, Z.G.; Jin, B. Function and mechanism of WRKY transcription factors in abiotic stress responses of plants. Plants 2020, 9, 1515. [Google Scholar] [CrossRef]
  19. Zhang, Y.; Du, P.; Xiong, F.; Zhang, X.J.; Song, H. WRKY Genes Improve Drought Tolerance in Arachis duranensis. J. Front. Plant Sci. 2022, 13, 910408. [Google Scholar] [CrossRef]
  20. Eulgem, T.; Rushton, P.J.; Robatzak, S.; Somssich, I.E. The WRKY superfamily of plant transcription factors. J. Trends Plant Sci. 2000, 5, 199–206. [Google Scholar] [CrossRef]
  21. Chen, L.G.; Song, Y.; Li, S.J.; Zhang, L.P.; Zou, C.S.; Yu, D.Q. The role of WRKY transcription factors in plant abiotic stresses. J. BBA-Gene Regul. Mech. 2012, 1819, 120–128. [Google Scholar] [CrossRef] [PubMed]
  22. Goyal, P.; Devi, R.; Verma, B.; Hussain, S.; Arora, P.; Tabassum, R.; Gupta, S. WRKY transcription factors: Evolution, regulation, and functional diversity in plants. J. Protoplasma 2023, 260, 331–348. [Google Scholar] [CrossRef] [PubMed]
  23. Zhu, Z.G.; Shi, J.L.; Cao, J.L.; He, M.Y.; Wang, Y.J. VpWRKY3 abiotic and abiotic stress-related transcription factor from the Chinese wild Vitis pseudoreticulata. J. Plant Cell Rep. 2012, 31, 2109–2120. [Google Scholar] [CrossRef] [PubMed]
  24. Zhao, L.; Zhang, W.; Song, Q.; Xuan, Y.; Li, K.; Cheng, L.; Qiao, H.; Wang, G.; Zhou, C. A WRKY transcription factor, TaWRKY40-D, promotes leaf senescence associated with jasmonic acid and abscisic acid pathways in wheat. J. Plant Biol. 2020, 22, 1072–1085. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, X.T.; Liu, J.; Lin, G.F.; Wang, A.; Wang, Z.H.; Lu, G.D. Overexpression of AtWRKY28 and AtWRKY75 in Arabidopsis enhances resistance to oxalic acid and Sclerotinia sclerotiorum. J. Plant Cell Rep. 2013, 32, 1589–1599. [Google Scholar] [CrossRef] [PubMed]
  26. Hu, Y.R.; Dong, Q.Y.; Yu, D.Q. Arabidopsis WRKY46 coordinates with WRKY70 and WRKY53 in basal resistance against the pathogen Pseudomonas syringae. J. Plant Sci. 2012, 185–186, 288–297. [Google Scholar] [CrossRef]
  27. Jones, J.D.G.; Dangl, J.L. The plant immune system. J. Nat. 2006, 444, 323–329. [Google Scholar] [CrossRef]
  28. Jones, J.D.G.; Vance, R.E.; Dangl, J.L. Intracellular innate immune surveillance devices in plants and animals. J. Sci. 2016, 354, 6395. [Google Scholar] [CrossRef]
  29. Gust, A.A.; Pruitt, R.; Nürnberger, T. Sensing danger: Key to activating plant immunity. J. Trends Plant Sci. 2017, 22, 779–791. [Google Scholar] [CrossRef]
  30. Saijo, Y.; Loo, E.P.-I.; Yasuda, S. Pattern recognition receptors and signaling in plant-microbe interactions. Plant J. 2018, 93, 592–613. [Google Scholar] [CrossRef]
  31. Shinde, B.A.; Dholakia, B.B.; Hussain, K.; Aharoni, A.; Giri, A.P.; Kamble, A.C. WRKY1 acts as a key component improving resistance against Alternaria solani in wild tomato, Solanum arcanum Peralta. Plant Biotechnol. J. 2018, 16, 1502–1513. [Google Scholar] [CrossRef] [PubMed]
  32. Zheng, L.L.; Qiu, B.L.; Su, L.L.; Wang, H.L.; Cui, X.M.; Ge, F.; Liu, D.Q. Panax notoginseng WRKY Transcription Factor 9 Is a Positive Regulator in Responding to Root Rot Pathogen Fusarium solani. J. Front. Plant Sci. 2022, 13, 930644. [Google Scholar] [CrossRef] [PubMed]
  33. Nakayama, A.; Fukushima, S.; Goto, S.; Matsushita, A.; Shimono, M.; Sugano, S.; Jiang, C.J.; Akagi, A.; Yamazaki, M.; Inoue, H.; et al. Genome-wide identification of WRKY45-regulated genes that mediate benzothiadiazole-induced defense responses in rice. J. BMC Plant Biol. 2013, 13, 150–160. [Google Scholar] [CrossRef] [PubMed]
  34. Fu, Y.Y.; Li, J.; Wu, H.; Jiang, S.J.; Zhu, Y.Y.; Liu, C.Y.; Xu, W.J.; Li, Q.; Yang, L.P. Analyses of Botrytis cinerea-responsive LrWRKY genes from Lilium regale reveal distinct roles of two LrWRKY transcription factors in mediating responses to B. cinerea. J. Plant Cell Rep. 2022, 41, 995–1012. [Google Scholar] [CrossRef]
  35. Wei, W.; Cui, M.Y.; Hu, Y.; Gao, K.; Xie, Y.G.; Jiang, Y.; Feng, J.Y. Ectopic expression of FvWRKY42 a WRKY transcription factor from the diploid woodland strawberry (Fragaria vesca), enhances resistance to powdery mildew, improves osmotic stress resistance, and increases abscisic acid sensitivity in Arabidopsis. J. Plant Sci. 2018, 275, 60–74. [Google Scholar] [CrossRef]
  36. Chen, C.H.; Chen, Z.X. Potentiation of developmentally regulated plant defense response by AtWRKY18, a pathogen-induced Arabidopsis transcription factor. J. Plant Physiol. 2002, 129, 706–716. [Google Scholar] [CrossRef]
  37. Zhu, J.; Zhong, S.F.; Guan, J.; Chen, W.; Yang, H.; Yang, H.; Chen, C.; Tan, F.Q.; Ren, T.H.; Li, Z.; et al. Genome-Wide Identification and Expression Analysis of WRKY Transcription Factors in Akebia trifoliata: A Bioinformatics Study. J. Genes 2022, 13, 1540. [Google Scholar] [CrossRef]
  38. Dong, J.X.; Chen, C.H.; Chen, Z.X. Expression profiles of the Arabidopsis WRKY gene superfamily during plant defense response. J. Plant Mol. Biol. 2003, 51, 21–37. [Google Scholar] [CrossRef]
  39. Yang, B.; Jiang, Y.Q.; Rahman, M.; Deyholos, M.; Kav, N. Identification and expression analysis of WRKY transcription factor genes in canola (Brassica napus L.) in response to fungal pathogens and hormone treatments. J. BMC Plant Biol. 2009, 9, 68. [Google Scholar] [CrossRef]
  40. Kalde, M.; Barth, M.; Somssich, I.E.; Lippok, B. Members of the Arabidopsis WRKY group III transcription factors are part of different plant defense signaling pathways. Mol. Plant-Microbe Interact. 2003, 16, 295–305. [Google Scholar] [CrossRef]
  41. Eulgem, T. Regulation of the Arabidopsis defense transcriptome. J. Trends Plant Sci. 2005, 10, 71–78. [Google Scholar] [CrossRef]
  42. Li, J.; Wang, J.; Wang, N.; Guo, X.Q.; Gao, Z. Ghwrky44, a WRKY transcription factor of cotton, mediates defense responses to pathogen infection in transgenic Nicotiana benthamiana. J. Plant Cell Tissue Organ Cult. 2015, 121, 127–140. [Google Scholar] [CrossRef]
  43. Liu, F.; Li, X.X.; Wang, M.R.; Wen, J.; Yi, B.; Shen, J.X.; Ma, C.Z.; Fu, T.D.; Tu, J.X. Interactions of Wrky15 and Wrky33 transcription factors and their roles in the resistance of Oilseed rape to Sclerotinia infection. Plant Biotechnol. J. 2018, 16, 911–925. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, J.; Tao, F.; Tian, W.; Guo, Z.F.; Chen, X.M.; Xu, X.M.; Shang, H.S.; Hu, X.P. The wheat WRKY transcription factors Tawrky49 and Tawrky62 confer differential high-temperature seedling-plant resistance to Puccinia striiformis F. sp. tritici. PLoS ONE 2017, 12, e0181963. [Google Scholar] [CrossRef] [PubMed]
  45. Cui, J.; Xu, P.; Meng, J.; Li, J.B.; Jiang, N.; Luan, Y.S. Transcriptome signatures of tomato leaf induced by phytophthora infestans and functional identification of transcription factor Spwrky3. J. Theor. Appl. Genet. 2018, 131, 787–800. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, Q.; Li, X.; Yan, S.; Yu, T.; Yang, J.Y.; Dong, J.F.; Zhang, S.H.; Zhao, J.L.; Yang, T.F.; Mao, X.X.; et al. Oswrky67 Positively regulates blast and bacteria blight resistance by direct activation of Pr genes in rice. J. BMC Plant Biol. 2018, 18, 257. [Google Scholar] [CrossRef]
  47. Luan, Q.; Chen, C.; Liu, M.Y.; Li, Q.; Wang, L.; Ren, Z.H. Cswrky50 mediates defense responses to pseudoperonospora cubensis infection in cucumis sativus. J. Plant Sci. 2019, 279, 59–69. [Google Scholar] [CrossRef]
  48. Shan, W.; Chen, J.Y.; Kuang, J.F.; Lu, W.J. Banana fruit nac transcription factor manac5 cooperates with mawrkys to enhance the expression of pathogenesis-related genes against Colletotrichum musae. J. Mol. Plant Pathol. 2016, 17, 330–338. [Google Scholar] [CrossRef]
  49. Hu, L.; Ye, M.; Li, R.; Lou, Y.G. Oswrky53, a versatile switch in regulating herbivore-induced defense responses in rice. J. Plant Signal. 2016, 11, e1169357. [Google Scholar] [CrossRef]
  50. Ma, Y.; Xue, H.; Zhang, L.; Zhang, F.; Ou, C.; Wang, F.; Zhang, Z. Involvement of auxin and brassinosteroid in dwarfism of autotetraploid apple (Malus × domestica). J. Sci. Rep. 2016, 6, 26719. [Google Scholar] [CrossRef]
  51. Rishmawi, L.; Juengst, C.; Schauss, A.C.; Schrader, A.; Hülskamp, M. Non-cell-autonomous regulation of root hair patterning genes by WRKY75 in Arabidopsis. J. Plant Physiol. 2014, 165, 186–195. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, D.; Chen, Q.Y.; Chen, W.W.; Liu, X.Y.; Xia, Y.; Guo, Q.G.; Jing, D.L.; Liang, G.L. A WRKY Transcription Factor, EjWRKY17, from Eriobotrya japonica Enhances Drought Tolerance in Transgenic Arabidopsis. Int. J. Mol. Sci. 2021, 22, 5593. [Google Scholar] [CrossRef] [PubMed]
  53. Jiang, Y.Z.; Guo, L.; Liu, R.; Jiao, B.; Zhao, X.; Ling, Z.Y.; Luo, K.M. Overexpression of Poplar PtrWRKY89 in Transgenic Arabidopsis Leads to a Reduction of Disease Resistance by Regulating Defense-Related Genes in Salicylate- and Jasmonate-Dependent Signaling. PLoS ONE 2016, 11, e0149137. [Google Scholar] [CrossRef]
  54. Hui, E.S.; Wang, P.F.; Lin, J.Y.; Zhao, C.Z.; Bi, Y.P.; Wang, X.J. Genome-wide identification and characterization of WRKY gene family in peanut. J. Front. Plant Sci. 2016, 7, 534. [Google Scholar]
  55. Rohit, J.; Shabir, H.W.; Balwant, S.; Abhishek, B.; Zahoor, A.D.; Ajaz, A.L.; Ashwani, P.; Sneh Lata, S.P. Transcription factors and plants response to drought stress: Current understanding and future directions. J. Front. Plant Sci. 2016, 7, 1029. [Google Scholar]
  56. Khoso, M.A.; Hussain, A.; Ritonga, F.N.; Ali, Q.; Channa, M.M.; Alshegaihi, R.M.; Meng, Q.L.; Ali, M.; Zaman, W.; Brohi, R.D.; et al. WRKY transcription factors (TFs): Molecular switches to regulate drought, temperature, and salinity stresses in plants. J. Front. Plant Sci. 2022, 13, 1039329. [Google Scholar] [CrossRef] [PubMed]
  57. Lee, F.C.; Yeap, W.C.; Appleton, D.R.; Ho, C.L.; Kulaveerasingam, H. Identification of drought-responsive Elaeis guineensis WRKY transcription factors with sensitivity to other abiotic stresses and hormone treatments. J. BMC Genom. 2022, 23, 164. [Google Scholar] [CrossRef]
  58. Grover, A.; Mittal, D.; Negi, M.; Dhruv, L. Generating high temperature tolerant transgenic plants: Achievements and challenges. J. Plant Sci. 2013, 205, 38–47. [Google Scholar] [CrossRef]
  59. Naohiko, O.; Hikaru, S.; Kazuo, S.; Kazuko, Y.S. Transcriptional regulatory network of plant heat stress response. J. Trends Plant Sci. 2017, 22, 53–65. [Google Scholar]
  60. Meher, J.; Lenka, S.; Sarkar, A.; Sarma, B.K. Transcriptional regulation of OsWRKY genes in response to individual and overlapped challenges of Magnaporthe oryzae and drought in indica genotypes of rice. J. Environ. Exp. Bot. 2023, 207, 105221. [Google Scholar] [CrossRef]
  61. Rizhsky, L.; Liang, H.R. The combined effect of drought stress and heat shock on gene expression in tobacco. J. Plant Physiol. 2002, 130, 1143–1151. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, X.; Xue, H.d.; Zhu, L.p.; Wang, H.Q.; Long, H.; Zhao, J.; Meng, F.N.; Liu, Y.F.; Ye, Y.; Luo, X.M.; et al. ERF49 mediates brassinosteroid regulation of heat stress tolerance in Arabidopsis thaliana. J. BMC Biol. 2022, 20, 254. [Google Scholar] [CrossRef] [PubMed]
  63. Jacob, P.; Hirt, H.; Bendahmane, A. The heat-shock protein/chaperone network and multiple stress resistance. Plant Biotechnol. J. 2017, 15, 405–414. [Google Scholar] [CrossRef] [PubMed]
  64. Dang, F.F.; Wang, Y.N.; Yu, L.; Eulgem, T.; Lai, Y.; Liu, Z.Q.; Wang, X.; Qiu, A.L.; Zhang, T.X.; Lin, J.; et al. CaWRKY40, a WRKY protein of pepper, plays an important role in the regulation of tolerance to heat stress and resistance to Ralstonia solanacearum infection. J. Plant Cell Environ. 2013, 36, 757–774. [Google Scholar] [CrossRef] [PubMed]
  65. Li, S.J.; Fu, Q.T.; Chen, L.G.; Huang, W.D.; Yu, D.Q. Arabidopsis thaliana WRKY25, WRKY26, and WRKY33 coordinate induction of plant thermotolerance. J. Planta 2011, 233, 1237–1252. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, J.; Tao, F.; An, F.; Zou, Y.P.; Tian, W.; Chen, X.M.; Xu, X.M.; Hu, X.P. Wheat transcription factor TaWRKY70 is positively involved in high-temperature seedling plant resistance to Puccinia striiformis f.sp.tritici. J. Mol. Plant Pathol. 2016, 18, 649–661. [Google Scholar] [CrossRef]
  67. Wu, Z.; Li, T.; Cao, X.; Zhang, D.; Teng, N.J. Lily WRKY factor LlWRKY22 promotes thermotolerance through autoactivation and activation of LlDREB2B. J. Hortic. Res. 2022, 9, uhac186. [Google Scholar] [CrossRef]
  68. Dou, L.L.; Guo, Y.N.; Ondati, E.; Pang, C.Y.; Wei, H.L.; Song, M.Z.; Fan, S.L.; Yu, S.X. Identification and expression analysis of group III WRKY transcription factors in cotton. J. Integr. Agric. 2016, 15, 2469–2480. [Google Scholar] [CrossRef]
  69. Guo, M.Y.; Yang, F.J.; Liu, C.X.; Zou, J.P.; Qi, Y.; Fotopoulos, V.; Lu, G.; Yu, J.Q.; Zhou, J. A single nucleotide polymorphism in the WRKY33 promoter is associated with the cold sensitivity in cultivated tomato. J. New Phytol. 2022, 236, 989–1005. [Google Scholar] [CrossRef]
  70. Yokotani, N.; Sato, Y.k.; Tanabe, S.; Chujo, T.; Shimizu, T.; Okada, K.; Yamane, H.; Shimono, M.; Sugano, S.; Takatsuji, H.; et al. WRKY76 is a rice transcriptional repressor playing opposite roles in blast disease resistance and cold stress tolerance. J. Exp. Bot. 2013, 64, 5085–5097. [Google Scholar] [CrossRef]
  71. Jiang, Y.Q.; Deyholos, M.K. Comprehensive transcriptional profiling of NaCl-stressed Arabidopsis roots reveals novel classes of responsive genes. J. BMC Plant Biol. 2006, 6, 25. [Google Scholar]
  72. Ma, Y.; Szostkiewicz, I.; Korte, A.; Moes, D.; Yang, Y.; Christmann, A.; Grill, E. Regulators of PP2C phosphatase activity function as abscisic acid sensors. Science 2009, 324, 1064–1068. [Google Scholar] [CrossRef] [PubMed]
  73. Cutler, S.R.; Rodriguez, P.L.; Finkelstein, R.R.; Abrams, S.R. Abscisic acid: Emergence of a core signaling network. Annu. Rev. Plant Biol. 2010, 61, 651–679. [Google Scholar] [CrossRef]
  74. Yan, H.R.; Jia, H.H.; Chen, X.B.; Hao, L.L.; An, H.L.; Guo, X.Q. The cotton WRKY transcription factor GhWRKY17 functions in drought and salt stress in transgenic Nicotiana benthamiana through ABA signaling and the modulation of reactive oxygen species production. J. Plant Cell Physiol. 2014, 55, 2060–2076. [Google Scholar] [CrossRef]
  75. Zhang, J.W.; Huang, D.Z.; Zhao, X.J.; Zhang, M.; Wang, Q.; Hou, X.Y.; Di, D.L.; Su, B.B.; Wang, S.K.; Sun, P. Drought-responsive WRKY transcription factor genes IgWRKY50 and IgWRKY32 from Iris germanica enhance drought resistance in transgenic Arabidopsis. J. Front. Plant Sci. 2022, 13, 983600. [Google Scholar] [CrossRef] [PubMed]
  76. Luo, X.; Bai, X.; Sun, X.L.; Zhu, D.; Liu, B.h.; Ji, W.; Cai, H.; Cao, L.; Wu, J.; Hu, M.R.; et al. Expression of wild soybean WRKY20 in Arabidopsis enhances drought tolerance and regulates ABA signaling. J. Exp. Bot. 2013, 8, 2155–2169. [Google Scholar] [CrossRef] [PubMed]
  77. Shi, W.Y.; Du, Y.T.; Ma, J.; Min, D.H.; Jin, L.G.; Chen, J.; Chen, M.; Zhou, Y.B.; Ma, Y.Z.; Xu, Z.S.; et al. The WRKY transcription factor GmWRKY12 confers drought and salt tolerance in soybean. Int. J. Mol. Sci. 2018, 19, 4087. [Google Scholar] [CrossRef]
  78. Li, Z.; Wang, N.N.; Gong, S.Y.; Lu, R.; Li, Y.; Li, X.B. Overexpression of a cotton (Gossypium hirsutum) WRKY gene, GhWRKY34, in Arabidopsis enhances salt-tolerance of the transgenic plants. J. Plant Physiol. Biochem. 2015, 96, 311–320. [Google Scholar]
  79. Chu, X.Q.; Wang, C.; Chen, X.B.; Lu, W.J.; Li, H.; Wang, X.L.; Hao, L.L.; Guo, X.Q. Correction: The Cotton WRKY Gene GhWRKY41 Positively Regulates Salt and Drought Stress Tolerance in Transgenic Nicotiana benthamiana. PLoS ONE 2016, 10, e0143022. [Google Scholar] [CrossRef]
  80. Hu, Q.; Ao, C.W.; Wang, X.R.; Wu, Y.F.; Du, X.Z. GhWRKY1-like, a WRKY transcription factor, mediates drought tolerance in Arabidopsis via modulating ABA biosynthesis. J. BMC Plant Biol. 2021, 21, 458. [Google Scholar] [CrossRef]
  81. Sun, S.; Chen, H.; Yang, Z.Q.; Lu, J.Y.; Wu, D.S.; Luo, Q.F.; Jia, J.; Tan, J.H. Identification of WRKY transcription factor family genes in Pinus massoniana Lamb. and their expression patterns and functions in response to drought stress. J. BMC Plant Biol. 2022, 22, 424. [Google Scholar] [CrossRef] [PubMed]
  82. Fu, Q.T.; Yu, D.Q. Expression profiles of AtWRKY25, AtWRKY26, and AtWRKY33 under abiotic stresses. J. Hered. 2010, 32, 848–856. [Google Scholar]
  83. Zhao, K.X.; Chu, S.S.; Zhang, X.D.; Wang, L.P.; Rono, J.K.; Yang, Z.M. AtWRKY21 negatively regulates tolerance to osmotic stress in Arabidopsis. J. Environ. Exp. Bot. 2020, 169, 103920. [Google Scholar] [CrossRef]
  84. Chen, S.Y.; Ding, Y.L.; Tian, H.N.; Wang, S.C.; Zhang, Y.L. WRKY54 and WRKY70 positively regulate SARD1 and CBP60g expression in plant immunity. J. Plant Signal. Behav. 2021, 16, 1932142. [Google Scholar] [CrossRef]
  85. Liu, Q.L.; Xu, K.D.; Pan, Y.Z.; Jiang, B.B.; Liu, G.L.; Yin, J.; Zhang, H.Q. Functional Analysis of a Novel Chrysanthemum WRKY Transcription Factor Gene Involved in Salt Tolerance. J. Plant Mol. Biol. Rep. 2014, 32, 282–289. [Google Scholar] [CrossRef]
  86. Yao, J.; Shen, Z.D.; Zhang, Y.L.; Wu, X.; Wang, J.H.; Sa, G.; Zhang, Y.H.; Zhang, H.L.; Deng, C.; Liu, J.; et al. Populus euphratica WRKY1 binds the promoter of H+-ATPase gene to enhance gene expression and salt tolerance. J. Exp. Bot. 2020, 71, 1527–1539. [Google Scholar] [CrossRef]
  87. Devaiah, B.N.; Karthikeyan, A.S.; Raghothama, K.G. WRKY75 transcription factor is a modulator of phosphate acquisition and root development in Arabidopsis. J. Plant Physiol. 2007, 143, 1789–1801. [Google Scholar] [CrossRef]
  88. Xu, L.; Jin, L.; Long, L.; Liu, L.L.; He, X.; Gao, W.; Zhu, L.F.; Zhang, X.L. Overexpression of GbWRKY1 positively regulates the Pi starvation response by alteration of auxin sensitivity in Arabidopsis. J. Plant Cell Rep. 2012, 31, 2177–2188. [Google Scholar] [CrossRef]
  89. Shen, N.; Hou, S.F.; Tu, G.Q.; Lan, W.Z.; Jing, Y.P. Transcription Factor WRKY33 Mediates the Phosphate Deficiency-Induced Remodeling of Root Architecture by Modulating Iron Homeostasis in Arabidopsis Roots. Int. J. Mol. Sci. 2021, 22, 9275. [Google Scholar] [CrossRef]
  90. Zhao, M.H.; Zhang, W.Z.; Ma, D.R.; Xu, Z.J.; Wang, J.Y.; Zhang, L.; Chen, W.F. Altered expression of transcription factor genes in rice flag leaf under low nitrogen stress. Rice Sci. 2012, 19, 100–107. [Google Scholar] [CrossRef]
  91. Li, J.; Xin, W.; Wang, W.P.; Zhao, S.J.; Xu, L.; Jiang, X.D.; Duan, Y.X.; Zheng, H.L.; Yang, L.M.; Liu, H.L.; et al. Mapping of Candidate Genes in Response to Low Nitrogen in Rice Seedlings. J. Rice 2022, 15, 51. [Google Scholar] [CrossRef] [PubMed]
  92. Yan, J.Y.; Li, C.X.; Sun, L.; Ren, J.Y.; Li, X.; Ding, Z.J.; Zheng, S.J. A WRKY transcription factor regulates Fe translocation under Fe deficiency. Plant Physiol. 2016, 171, 2017–2027. [Google Scholar] [CrossRef] [PubMed]
  93. Ichiro, K.; Yoko, I.; Masami, Y.H.; Toru, F. WRKY6 is involved in the response to boron deficiency in Arabidopsis thaliana. Physiol. Plant 2010, 139, 80–92. [Google Scholar]
  94. Ding, Z.J.; Yan, J.Y.; Xu, X.Y.; Li, G.X.; Zheng, S.J. WRKY46 functions as a transcriptional repressor of ALMT1, regulating aluminum-induced malate secretion in Arabidopsis. Plant J. Cell Mol. Biol. 2013, 76, 825–835. [Google Scholar] [CrossRef]
  95. Wu, X.L.; Chen, Q.; Chen, L.L.; Tian, F.F.; Chen, X.X.; Han, C.Y.; Mi, J.X.; Lin, X.Y.; Wan, X.Q.; Jiang, B.B.; et al. A WRKY transcription factor, PyWRKY75, enhanced cadmium accumulation and tolerance in poplar. J. Ecotoxicol. Environ. Saf. 2022, 239, 113630. [Google Scholar] [CrossRef]
  96. Dai, X.; Wang, Y.; Zhang, W.H. OsWRKY74, a WRKY transcription factor, modulates tolerance to phosphate starvation in rice. J. Exp. Bot. 2015, 67, 947–960. [Google Scholar] [CrossRef]
  97. Muhammad, Z.H.; Sabir, H.; Pia Muhammad, A.R.; Mutahar, I.; Muhammad, I.; Tanvir, S.; Maryam, F.; Shahbaz, A.K.; Imran, K.; Muhammad, S.; et al. Bentonite and Biochar Mitigate Pb Toxicity in Pisum sativum by Reducing Plant Oxidative Stress and Pb Translocation. J. Plants 2019, 8, 571. [Google Scholar]
  98. Tang, H.; Bi, H.; Liu, B.; Lou, S.L.; Song, Y.; Tong, S.F.; Chen, N.N.; Jiang, Y.Z.; Liu, J.Q.; Liu, H.H. WRKY33 interacts with WRKY12 protein to up-regulate RAP2.2 during submergence-induced hypoxia response in Arabidopsis thaliana. New Phytol. 2020, 229, 106–125. [Google Scholar] [CrossRef]
  99. Miao, Y.; Laun, T.M.; Smykowski, A.; Zentgraf, U. Arabidopsis MEKKI can take a shortcut: It can directly interact with senescence-related WRKY53 transcription factor on the protein level and can bind to its promoter. J. Plant Mol. Biol. 2007, 65, 63–76. [Google Scholar] [CrossRef]
  100. Sun, Y.; Yu, D. Activated expression of AtWRKY53 negatively regulates drought tolerance by mediating stomatal movement. J. Plant Cell Rep. 2015, 34, 1295–1306. [Google Scholar] [CrossRef]
  101. Jia, H.; Wang, C.; Wang, F.; Liu, S.C.; Li, G.L.; Guo, X.Q. GhWRKY68 reduces resistance to salt and drought in transgenic Nicotiana benthamiana. PLoS ONE 2017, 10. [Google Scholar] [CrossRef]
  102. Liu, W.; Liang, X.Q.; Cai, W.J.; Wang, H.; Liu, X.; Cheng, L.F.; Song, P.H.; Luo, G.J.; Han, D.G. Isolation and Functional Analysis of VvWRKY28 a Vitis vinifera WRKY Transcription Factor Gene, with Functions in Tolerance to Cold and Salt Stress in Transgenic Arabidopsis thaliana. Int. J. Mol. Sci. 2022, 23, 13418. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, Y.; Zhu, R.Y.; Shi, M.; Huang, Q.K.; Zhang, S.W.; Kai, G.Y.; Guo, S.L. Genome-Wide Identification and Comparative Analysis of WRKY Transcription Factors Related to Momilactone Biosynthesis in Calohypnum plumiforme. J. Front. Ecol. Evol. 2022, 9, 809729. [Google Scholar] [CrossRef]
  104. Xiao, Y.; Feng, J.X.; Li, Q.; Zhou, Y.Y.; Bu, Q.T.; Zhou, J.H.; Tan, H.X.; Yang, Y.B.; Zhang, L.; Chen, W.S. IiWRKY34 positively regulates yield, lignan biosynthesis, and stress tolerance in Isatis indigotica. J. Acta Pharm. Sin. B 2019, 10, 2417–2432. [Google Scholar] [CrossRef]
  105. Fu, J.Y.; Liu, Q.; Wang, C.; Liang, J.; Liu, L.J.; Wang, Q. ZmWRKY79 positively regulates maize phytoalexin biosynthetic gene expression and is involved in stress response. J. Exp. Bot. 2018, 69, 497–510. [Google Scholar] [CrossRef]
  106. Kato, N.; Dubouzet, E.; Kokabu, Y.; Yoshida, S.; Taniguchi, Y.; Dubouzet, J.G.; Yazaki, K.; Sato, F. Identification of a WRKY protein as a transcriptional regulator of benzylisoquinoline alkaloid biosynthesis in Coptis japonica. J. Plant Cell Physiol. 2007, 48, 8–18. [Google Scholar] [CrossRef]
  107. Suttipanta, N.; Pattanaik, S.; Kulshrestha, M.; Patra, B.; Singh Sanjay, K.; Yuan, L. The transcription factor CrWRKY1 positively regulates the terpenoid indole alkaloid biosynthesis in Catharanthus roseus. J. Plant Physiol. 2011, 157, 2081–2093. [Google Scholar] [CrossRef]
  108. Wang, C.; Hao, X.L.; Wang, Y.; Itay, M.; Zhou, W.; Zhou, W.; Zhou, Z.G.; Kai, G.Y. Identification of WRKY transcription factors involved in regulating the biosynthesis of the anticancer drug camptothecin in Ophiorrhiza pumila. Hortic. Res. 2022, 9, uhac099. [Google Scholar] [CrossRef]
  109. Spyropoulou, E.A.; Haring, M.A.; Schuurink, R.C. RNA sequencing on Solanum Lycopersicum trichomes identifies transcription factors that activate terpene synthase promoters. J. BMC Genom. 2014, 15, 402. [Google Scholar] [CrossRef]
  110. Ma, D.M.; Pu, G.B.; Lei, C.Y.; Ma, L.Q.; Wang, H.H.; Guo, Y.W.; Chen, J.L.; Du, Z.G.; Wang, H.; Li, G.F.; et al. Isolation and characterization of AaWRKY1, an Artemisia annua transcription factor that regulates the amorpha-4,11-diene synthase gene, a key gene of artemisinin biosynthesis. J. Plant Cell Physiol. 2009, 50, 2146–2161. [Google Scholar] [CrossRef]
  111. Xu, Y.H.; Wang, J.W.; Wang, S.; Wang, J.Y.; Chen, X.Y. Characterization of GaWRKY1, a cotton transcription factor that regulates the sesquiterpene synthase gene (+)-delta-cadinene synthase-A. J. Plant Physiol. 2004, 135, 507–515. [Google Scholar] [CrossRef] [PubMed]
  112. Duan, S.W.; Wang, J.J.; Gao, C.H.; Jin, C.Y.; Li, D.; Peng, D.S.; Du, G.M.; Li, Y.Q.; Chen, M.X. Functional characterization of a heterologously expressed Brassica napus WRKY41-1 transcription factor in regulating anthocyanin biosynthesis in Arabidopsis thaliana. J. Plant Sci. 2018, 268, 47–53. [Google Scholar] [CrossRef] [PubMed]
  113. Qiu, D.Y.; Xiao, J.; Xie, W.B.; Cheng, H.T.; Li, X.H.; Wang, S.P. Exploring transcriptional signaling mediated by OsWRKY13, a potential regulator of multiple physiological processes in rice. J. BMC Plant Biol. 2009, 9, 74. [Google Scholar]
  114. Wang, L.J.; Guo, D.H.; Zhao, G.D.; Wang, J.Y.; Zhang, S.X.; Wang, C.; Guo, X.Q. Group ΙΙc WRKY transcription factors regulate cotton resistance to Fusarium oxysporum by promoting GhMKK2-mediated flavonoid biosynthesis. New Phytol. 2022, 236, 249–265. [Google Scholar] [CrossRef] [PubMed]
  115. Zentgraf, U.; Laun, T.; Miao, Y. The complex regulation of WRKY53 during leaf senescence of Arabidopsis thaliana. Eur. J. Cell Biol. 2010, 89, 133–137. [Google Scholar] [CrossRef]
  116. Turck, F.; Zhou, A.; Somssich, I.E. Stimulus-dependent promoter-specific binding of transcription factor WRKY1 to Its native promoter and the defense-related gene PcPR1-1 in Parsley. J. Plant Cell 2004, 16, 2573–2585. [Google Scholar] [CrossRef]
  117. Xu, L.; Chen, X.; Zhu, D.Z.; Zong, X.J.; Wei, H.R.; Wang, J.W.; Tan, Y.; Liu, Q.Z. Cloning of PcWRKY3 gene and its expression in response to abiotic stress factors in cherry rootstocks. J. Acta Hortic. 2019, 93–98. [Google Scholar] [CrossRef]
  118. Jia, C.H.; Wang, Z.; Wang, J.Y.; Miao, H.X.; Zhang, J.B.; Xu, B.Y.; Liu, J.H.; Jin, Z.Q.; Liu, J.H. Genome-Wide Analysis of the Banana WRKY Transcription Factor Gene Family Closely Related to Fruit Ripening and Stress. J. Plants 2022, 11, 662. [Google Scholar] [CrossRef]
  119. Grzechowiak, M.; Ruszkowska, A.; Sliwiak, J.; Urbanowicz, A.; Jaskolski, M.; Ruszkowski, M. New aspects of DNA recognition by group II WRKY transcription factor revealed by structural and functional study of AtWRKY18 DNA binding domain. Int. J. Biol. Macromol. 2022, 213, 589–601. [Google Scholar] [CrossRef]
  120. Maeo, K.S.; Kojima-Suzuki, H.A.; Nakamura, K. Role of conserved residues of the WRKY domain in the DNA-binding of tobacco WRKY family proteins. J. Biosci. Biotechnol. Biochem. 2001, 65, 2428–2436. [Google Scholar] [CrossRef]
  121. Peizhen, Y.; Chen, C.H.; Wang, Z.P.; Fan, B.F.; Chen, Z.X. A pathogen-and salicylic acid-induced WRKY DNA-binding activity recognizes the elicitor response element of the tobacco class I chitinase gene promoter. Plant J. 1999, 18, 141–149. [Google Scholar]
  122. Zhang, H.; Li, D.; Wang, M.; Liu, W.; Teng, W.J.; Cheng, B.P.; Huang, Q.; Wang, M.; Song, W.W.; Dong, S.M.; et al. The Nicotiana benthamiana mitogen-activated protein kinase cascade and WRKY transcription factor participate in Nep1(Mo)-triggered plant responses. Mol. Plant-Microbe Interact. 2012, 25, 1639. [Google Scholar] [CrossRef]
  123. Li, L.Q.; Huang, L.P.; Pan, G.; Liu, L.; Wang, X.Y.; Lu, L.M.; He, Y.K.; Wang, X.C.; Dai, S.J. Identifying the Genes Regulated by AtWRKY6 Using Comparative Transcript and Proteomic Analysis under Phosphorus Deficiency. Int. J. Mol. Sci. 2017, 18, 1046. [Google Scholar] [CrossRef] [PubMed]
  124. Su, T.; Xu, Q.; Zhang, F.C.; Chen, Y.; Li, L.Q.; Wu, W.H.; Chen, Y.F. WRKY42 modulates phosphate homeostasis through regulating phosphate translocation and acquisition in Arabidopsis. J. Plant Physiol. 2015, 167, 1579–1591. [Google Scholar] [CrossRef] [PubMed]
  125. Hu, Y.R.; Chen, L.G.; Wang, H.P.; Zhang, L.P.; Wang, F.; Yu, D.Q. Arabidopsis transcription factor WRKY8 functions antagonistically with its interacting partner VQ9 to modulate salinity stress tolerance. Plant J. 2013, 74, 730–745. [Google Scholar] [CrossRef] [PubMed]
  126. GarcýaLaynes, S.; HerreraValencia, V.A.; TamayoTorres, L.G.; LimonesBriones, V.; BarredoPool, F.A.; BaasEspinola, F.M.; AlpucheSolís, A.G.; PuchHau, C.; PerazaEcheverria, S. The Banana MaWRKY18, MaWRKY45, MaWRKY60, and MaWRKY70 Genes Encode Functional Transcription Factors and Display Differential Expression in Response to Defense Phytohormones. Genes 2022, 13, 1891. [Google Scholar] [CrossRef]
  127. He, Y.G.; Zhu, M.H.; Li, Z.H.; Jiang, S.; He, Z.J.; Xu, S.; Chen, X.S.; Hu, Z.L.; Zhang, Z.H. IPA1 Negatively Regulates Early Rice Seedling Development by Interfering with Starch Metabolism via the GA and WRKY Pathways. Int. J. Mol. Sci. 2021, 22, 6605. [Google Scholar] [CrossRef]
  128. Liu, X.T.; Zhou, X.F.; Li, D.D.; Hong, B.; Gao, J.P.; Zhang, Z. Rose WRKY13 promotes disease protection to Botrytis by enhancing cytokinin content and reducing abscisic acid signaling. J. Plant Physiol. 2022, 191, 679–693. [Google Scholar] [CrossRef]
  129. Jiang, Y.J.; Yu, D.Q. The WRKY57 Transcription Factor Affects the Expression of Jasmonate ZIM-Domain Genes Transcriptionally to Compromise Botrytis cinerea Resistance. J. Plant Physiol. 2016, 171, 2771–2782. [Google Scholar] [CrossRef]
  130. Wang, T.J.; Huang, S.Z.; Zhang, A.; Guo, P.; Liu, Y.T.; Xu, C.M.; Cong, W.X.; Liu, B.; Xu, Z.Y. JMJ17–WRKY40 and HY5–ABI5 modules regulate the expression of ABA-responsive genes in Arabidopsis. J. New Phytol. 2021, 230, 567–584. [Google Scholar] [CrossRef]
  131. Yu, F.F.; Huaxia, Y.F.; Lu, W.J.; Wu, C.G.; Cao, X.C.; Guo, X.Q. GhWRKY15, a member of the WRKY transcription factor family identified from cotton (Gossypium hirsutum L.), is involved in disease resistance and plant development. J. BMC Plant Biol. 2012, 12, 144. [Google Scholar]
  132. Ay, N.; Irmler, K.; Fischer, A.; Uhlemann, R.; Reuter, G.; Humbeck, K. Epigenetic programming via histone methylation at WRKY53 controls leaf senescence in Arabidopsis thaliana: Epigenetic control of leaf senescence. Plant J. 2009, 58, 333–346. [Google Scholar] [CrossRef] [PubMed]
  133. Zhang, Y.; Yu, H.; Yang, X.; Li, Q.; Ling, J.; Wang, H.; Gu, X.F.; Huang, S.W.; Jiang, W.J. CsWRKY46 a WRKY transcription factor from cucumber, confers cold resistance in transgenic plants by regulating a set of cold-stress responsive genes in an ABA-dependent manner. J. Plant Physiol. Biochem. 2016, 108, 478–487. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, J.L.; Liu, B.; Song, Y.; Chen, Y.; Fu, J.; Liu, J.Q.; Ma, T.; Xi, Z.X.; Liu, H.H. Genome-wide (ChIP-seq) identification of target genes regulated by WRKY33 during submergence stress in Arabidopsis. J. BMC Genom. Data 2021, 22, 16. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, J.; Fu, X.J.; Zhang, S.; Chen, G.; Li, S.J.; Shangguan, T.W.; Zheng, Y.T.; Xu, F.; Chen, Z.H.; Xu, S.C. Evolutionary and Regulatory Pattern Analysis of Soybean Ca2+ ATPases for Abiotic Stress Tolerance. J. Front. Plant Sci. 2022, 13, 898256. [Google Scholar] [CrossRef] [PubMed]
  136. Köster, P.; DeFalco, T.A.; Zipfel, C. Ca2+ signals in plant immunity. EMBO J. 2022, 41, e110741. [Google Scholar] [CrossRef]
Figure 1. The WRKY domain. The WRKY motif as well as the cysteine and histidine residues that form zinc fingers are shown. Four β-chains are indicated by arrows. Group Ⅰ NT and CT represent the N-terminal and C-terminal WRKY domains of Group I WRKY proteins.
Figure 1. The WRKY domain. The WRKY motif as well as the cysteine and histidine residues that form zinc fingers are shown. Four β-chains are indicated by arrows. Group Ⅰ NT and CT represent the N-terminal and C-terminal WRKY domains of Group I WRKY proteins.
Cimb 45 00187 g001
Figure 2. Arabidopsis WRKY gene family phylogenetic tree.
Figure 2. Arabidopsis WRKY gene family phylogenetic tree.
Cimb 45 00187 g002
Figure 3. Mechanism underlying the regulatory effects of WRKY transcription factors on plant stress responses.
Figure 3. Mechanism underlying the regulatory effects of WRKY transcription factors on plant stress responses.
Cimb 45 00187 g003
Table 1. Regulation of WRKY transcription factors in biotic stress responses.
Table 1. Regulation of WRKY transcription factors in biotic stress responses.
Plant SpeciesGeneSpecies of Pathogenic Bacteria/FungiWay of ParticipationReferences
Wild tomato (Solanum arcanum Peralta)SaWRKY1Alternaria solaniInteraction pathway between plant pathogens: ETI[31]
Panax notoginsengPnWRKY9Fusarium solaniMethyl jasmonate (MeJA) signal transduction pathway[32]
Rice (Oryza sativa)OsWRKY45Magnaporthe oryzaeInteraction pathway between plant pathogens: ETI[33]
Lilium regale (Lilium browniivar)LrWRKY39Botrytis cinereaInteraction pathway between plant pathogens: ETI[34]
LrWRKY41aInteraction pathway between plant pathogens: ETI
Diploid woodland Strawberry (Fragaria vesca)FvWRKY42Sphaerotheca aphanisInteraction pathway between plant pathogens: ETI[35]
Arabidopsis (Arabidopsis thaliana)AtWRKY18Pseudomonas syringaeInteraction pathway between plant pathogens: ETI[36]
Three-leaf akebia (Akebia trifoliata)AkWRKY24 Interaction pathway between plant pathogens: ETI[37]
Arabidopsis (Arabidopsis thaliana)AtWRKY Interaction pathway between plant pathogens: ETI; salicylic acid (SA) signal transduction pathway[38,39,40,41]
Cotton (Gossypium hirsutum)GhWRKY44Ralstonia solanacerumInteraction pathway between plant pathogens: PTI, ETI; salicylic acid (SA) signal transduction pathway[42]
Cotton (Gossypium hirsutum)GhWRKY44Rhizoctonia solaniInteraction pathway between plant pathogens: PTI, ETI; salicylic acid (SA) signal transduction pathway[42]
Oilseed rape (Brassica napus)BnWRKY15, BnWRKY33Sclerotinia sclerotiorumInteraction pathway between plant pathogens: ETI; salicylic acid (SA) and jasmonic acid (JA) signal transduction pathways[43]
Wheat (Triticum aestivum)TaWRKY49, TaWRKY62Puccinia striiformisInteraction pathway between plant pathogens: PTI; salicylic acid (SA), jasmonic acid (JA), ethylene (ET) signal transduction pathway[44]
Tomato (Solanum pimpinellifolium)SpWRKY3Phytophthora infestansInteraction pathway between plant pathogens: ETI[45]
Rice (Oryza sativa)OsWRKY67Bacteria blightInteraction pathway between plant pathogens: ETI; salicylic acid (SA) signal transduction pathway[46]
Cucumber (Cumis sativus)CsWRKY50Pseudoperonospora ubensisInteraction pathway between plant pathogens: PTI, ETI; salicylic acid (SA) and jasmonic acid (JA) signal transduction pathways[47]
Banana (Musa acuminata)MaNAC5/MaWRKY1/MaWRKY2Colletotrchum musaeInteraction pathway between plant pathogens: ETI[48]
Rice (Oryza sativa)OsWRKY53Chilo suppressalisInteraction pathway between plant pathogens: PTI; ethylene (ET) signal transduction pathway[49]
Arabidopsis (Arabidopsis thaliana)WRKY75 Interaction pathway between plant pathogens: PTI; ethylene (ET) signal transduction pathway[50,51]
Loquat (Eriobotrya japonica)EjWRKY17 Interaction pathway between plant pathogens: ETI; abscisic-acid (ABA) signal transduction pathway[52]
Poplar (Populus trichocarpa)PtrWRY89 Interaction pathway between plant pathogens: ETI; salicylic acid (SA) signal transduction pathway[53]
Peanut (Arachis hypogaea)AhWRY1 Salicylic acid (SA), and jasmonic acid (JA) signal transduction pathway[54]
AhWRY7
AhWRY8
AhWRY12
AhWRY13
Table 3. WRKY transcription factors that induce oxidative stress responses during an exposure to other stresses.
Table 3. WRKY transcription factors that induce oxidative stress responses during an exposure to other stresses.
Type of StressSpeciesGeneSignal Transduction PathwayReferences
Apx1Arabidopsis (Arabidopsis thaliana)WRKY25Zinc-finger protein Zat12[97]
FloodingArabidopsis (Arabidopsis thaliana)WRKY33, WRKY12Ethylene response factor VII gene RAP2.2[98]
Arabidopsis (Arabidopsis thaliana)WRKY53MEKK1[99]
starch degradation[100]
AtWRKY8MAPKKα-MEK2-WIPK signaling cascade downstream gene interaction[101]
Drought and saltCotton (Gossypium hirsutum)GhWRKY68Accumulate more H2O2 and O2-[101]
GhWRKY17Increased reactive oxygen species (ROS) levels[74]
Low temperature and saltGrape (Vitis vinifera)VvWRKY28Superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT) synthesis pathways[102]
Table 4. Regulatory effects of WRKY transcription factors on plant secondary metabolism.
Table 4. Regulatory effects of WRKY transcription factors on plant secondary metabolism.
SpeciesGeneSecondary MetabolitesRegulation ModeReferences
Isatis indigotica (Isatis tinctoria)IiWRKY34LigninPositive regulation[104]
Maize (Zea mays)ZmWRKY79PhytohormonePositive regulation[105]
Coptis japonicaCjWRKY1AlkaloidPositive regulation[106]
Catharanthus roseusCrWRKY1Terpenoid indole alkaloidsPositive regulation[107]
Dwarf lilyturf root (Ophiorhiz pumila)OpWRKY6CamptothecinNegative regulation[108]
Tomato (Solanum lycopersicum)SlWRKY71TerpenoidsPositive regulation[109]
Artemisia annuaAaWRKY1ArtemisininPositive regulation[110]
Cotton (Gossypium hirsutum)GaWRKY1Sesquiterpene aldehydePositive regulation[111]
Oilseed rape (Brassica napus)BnWRKY41-1CyanidinNegative regulation[112]
Rice (Oryza sativa)OsWRKY13FlavonoidsPositive regulation[113]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, H.; Cheng, X.; Yin, D.; Chen, D.; Luo, C.; Liu, H.; Huang, C. Advances in the Research on Plant WRKY Transcription Factors Responsive to External Stresses. Curr. Issues Mol. Biol. 2023, 45, 2861-2880. https://doi.org/10.3390/cimb45040187

AMA Style

Wang H, Cheng X, Yin D, Chen D, Luo C, Liu H, Huang C. Advances in the Research on Plant WRKY Transcription Factors Responsive to External Stresses. Current Issues in Molecular Biology. 2023; 45(4):2861-2880. https://doi.org/10.3390/cimb45040187

Chicago/Turabian Style

Wang, Hongli, Xi Cheng, Dongmei Yin, Dongliang Chen, Chang Luo, Hua Liu, and Conglin Huang. 2023. "Advances in the Research on Plant WRKY Transcription Factors Responsive to External Stresses" Current Issues in Molecular Biology 45, no. 4: 2861-2880. https://doi.org/10.3390/cimb45040187

Article Metrics

Back to TopTop